Cyclopropenium ion

Last updated
The cyclopropenium cation CyclopropeniumDeloc.png
The cyclopropenium cation

The cyclopropenium ion is the cation with the formula C
3
H+
3
. It has attracted attention as the smallest example of an aromatic cation. Its salts have been isolated, and many derivatives have been characterized by X-ray crystallography. [1] The cation and some simple derivatives have been identified in the atmosphere of the Saturnian moon Titan. [2]

Contents

Bonding

With two π electrons, the cyclopropenium cation class obeys Hückel’s rules of aromaticity for 4n + 2 electrons since, in this case, n = 0. Consistent with this prediction, the C3H3 core is planar and the C–C bonds are equivalent. In the case of the cation in [C3(SiMe3)3]+SbCl
6
, [3] the ring C–C distances range from 1.374(2) to 1.392(2) Å.

Structure of the salt [C3(SiMe3)3]SbCl
6 NEZKOS01.png
Structure of the salt [C3(SiMe3)3]SbCl
6

Syntheses

Salts of many cyclopropenyl cations have been characterized. Their stability varies according to the steric and inductive effects of the substituents.

Salts of triphenylcyclopropenium were first reported by Ronald Breslow in 1957. The salt was prepared in two steps starting with the reaction of phenyldiazoacetonitrile with diphenylacetylene to yield 1,2,3-triphenyl-2-cyclopropene nitrile. Treatment of this with boron trifluoride yielded [C3Ph3]BF4. [4] [5] [6]

Cyclopropenium synthesis3.png

The parent cation, [C3H3]+, was reported as its hexachloroantimonate (SbCl
6
) salt in 1970. [7] It is indefinitely stable at −20 °C.

Trichlorocyclopropenium salts are generated by chloride abstraction from tetrachlorocyclopropene: [8]

C3Cl4 + AlCl3 → [C3Cl3]+AlCl
4

Tetrachlorocyclopropene can be converted to tris(tert-butyldimethylsilyl)cyclopropene. Hydride abstraction with nitrosonium tetrafluoroborate yields the trisilyl-substituted cyclopropenium cation. [9]

Another synthesis of Cyclopropenium 2.png

Amino-substituted cyclopropenium salts are particularly stable. [10] [11] Calicene is an unusual derivative featuring cyclopropenium linked to a cyclopentadienide.

Calicene features a cyclopropenium ring. CaliceneDeloc.png
Calicene features a cyclopropenium ring.

Reactions

Organic chemistry

Chloride salts of cyclopropenium esters are intermediates in the use of dichlorocyclopropenes for the conversion of carboxylic acids to acid chlorides: [12]

Acidchloride2.jpg

Related cyclopropenium cations are produced in the regeneration of the 1,1-dichlorocyclopropenes from the cyclopropenones.

The cyclopropenium chlorides have been applied to peptide bond formation. [12] For example, in the figure below, reacting a boc-protected amino acid with an unprotected amino acid in the presence of the cyclopropenium ion allows the formation of a peptide bond via acid chloride formation followed by nucleophilic substitution with the unprotected amino acid.

Peptidecyclo2.jpg

This method of mildly generating acid chlorides can also be useful for linking alpha-anomeric sugars. [13] After using the cyclopropenium ion to form the chloride at the anomeric carbon, the compound is iodinated with tetrabutylammonium iodide. This iodine can thereafter be substituted by any ROH group to quickly undergo alpha-selective linkage of sugars.

Alphaanomericlinkagecyclo3.jpg

Additionally, some synthetic routes make use of cyclopropenium ring openings yielding an allylcarbene cation. The linear degradation product yields both a nucleophilic and electrophilic carbon centers. [14]

Bimodal Ring Opening of Cyclopropenium Ion.png

Organometallic compounds

Structure of Ph3C3Co(CO)3 viewed down the C3 symmetry axis. TPCPCQ.png
Structure of Ph3C3Co(CO)3 viewed down the C3 symmetry axis.

Many complexes are known with cyclopropenium ligands. Examples include [M(C3Ph3)(PPh3)2]+ (M = Ni, Pd, Pt) and Co(C3Ph3)(CO)3. Such compounds are prepared by reaction of cyclopropenium salts with low valent metal complexes. [15]

As polyelectrolytes

Because many substituted derivatives are known, cyclopropenium salts have attracted attention as possible polyelectrolytes, relevant to technologies such as desalination and fuel cells. The tris(dialkylamino)cyclopropenium salts have been particularly evaluated because of their high stability. [16]

See also

Related Research Articles

<span class="mw-page-title-main">Amine</span> Chemical compounds and groups containing nitrogen with a lone pair (:N)

In chemistry, amines are compounds and functional groups that contain a basic nitrogen atom with a lone pair. Amines are formally derivatives of ammonia, wherein one or more hydrogen atoms have been replaced by a substituent such as an alkyl or aryl group. Important amines include amino acids, biogenic amines, trimethylamine, and aniline. Inorganic derivatives of ammonia are also called amines, such as monochloramine.

The Sandmeyer reaction is a chemical reaction used to synthesize aryl halides from aryl diazonium salts using copper salts as reagents or catalysts. It is an example of a radical-nucleophilic aromatic substitution. The Sandmeyer reaction provides a method through which one can perform unique transformations on benzene, such as halogenation, cyanation, trifluoromethylation, and hydroxylation.

<i>N</i>-Bromosuccinimide Molecule

N-Bromosuccinimide or NBS is a chemical reagent used in radical substitution, electrophilic addition, and electrophilic substitution reactions in organic chemistry. NBS can be a convenient source of Br, the bromine radical.

<span class="mw-page-title-main">Diazonium compound</span> Group of organonitrogen compounds

Diazonium compounds or diazonium salts are a group of organic compounds sharing a common functional group [R−N+≡N]X where R can be any organic group, such as an alkyl or an aryl, and X is an inorganic or organic anion, such as a halide.

<span class="mw-page-title-main">Curtius rearrangement</span> Chemical reaction

The Curtius rearrangement, first defined by Theodor Curtius in 1885, is the thermal decomposition of an acyl azide to an isocyanate with loss of nitrogen gas. The isocyanate then undergoes attack by a variety of nucleophiles such as water, alcohols and amines, to yield a primary amine, carbamate or urea derivative respectively. Several reviews have been published.

<span class="mw-page-title-main">Persistent carbene</span> Type of carbene demonstrating particular stability

A persistent carbene is an organic molecule whose natural resonance structure has a carbon atom with incomplete octet, but does not exhibit the tremendous instability typically associated with such moieties. The best-known examples and by far largest subgroup are the N-heterocyclic carbenes (NHC), in which nitrogen atoms flank the formal carbene.

<span class="mw-page-title-main">Carbenium ion</span> Class of ions

A carbenium ion is a positive ion with the structure RR′R″C+, that is, a chemical species with carbon atom having three covalent bonds, and it bears a +1 formal charge. But IUPAC confuses coordination number with valence, incorrectly considering carbon in carbenium as trivalent.

<span class="mw-page-title-main">Petasis reaction</span>

The Petasis reaction is the multi-component reaction of an amine, a carbonyl, and a vinyl- or aryl-boronic acid to form substituted amines.

<span class="mw-page-title-main">Dakin oxidation</span> Organic redox reaction that converts hydroxyphenyl aldehydes or ketones into benzenediols

The Dakin oxidation (or Dakin reaction) is an organic redox reaction in which an ortho- or para-hydroxylated phenyl aldehyde (2-hydroxybenzaldehyde or 4-hydroxybenzaldehyde) or ketone reacts with hydrogen peroxide (H2O2) in base to form a benzenediol and a carboxylate. Overall, the carbonyl group is oxidised, whereas the H2O2 is reduced.

Pyrylium is a cation with formula C5H5O+, consisting of a six-membered ring of five carbon atoms, each with one hydrogen atom, and one positively charged oxygen atom. The bonds in the ring are conjugated as in benzene, giving it an aromatic character. In particular, because of the positive charge, the oxygen atom is trivalent. Pyrilium is a mono-cyclic and heterocyclic compound, one of the oxonium ions.

The Stetter reaction is a reaction used in organic chemistry to form carbon-carbon bonds through a 1,4-addition reaction utilizing a nucleophilic catalyst. While the related 1,2-addition reaction, the benzoin condensation, was known since the 1830s, the Stetter reaction was not reported until 1973 by Dr. Hermann Stetter. The reaction provides synthetically useful 1,4-dicarbonyl compounds and related derivatives from aldehydes and Michael acceptors. Unlike 1,3-dicarbonyls, which are easily accessed through the Claisen condensation, or 1,5-dicarbonyls, which are commonly made using a Michael reaction, 1,4-dicarbonyls are challenging substrates to synthesize, yet are valuable starting materials for several organic transformations, including the Paal–Knorr synthesis of furans and pyrroles. Traditionally utilized catalysts for the Stetter reaction are thiazolium salts and cyanide anion, but more recent work toward the asymmetric Stetter reaction has found triazolium salts to be effective. The Stetter reaction is an example of umpolung chemistry, as the inherent polarity of the aldehyde is reversed by the addition of the catalyst to the aldehyde, rendering the carbon center nucleophilic rather than electrophilic.

<span class="mw-page-title-main">Scholl reaction</span>

The Scholl reaction is a coupling reaction between two arene compounds with the aid of a Lewis acid and a protic acid. It is named after its discoverer, Roland Scholl, a Swiss chemist.

The Stieglitz rearrangement is a rearrangement reaction in organic chemistry which is named after the American chemist Julius Stieglitz (1867–1937) and was first investigated by him and Paul Nicholas Leech in 1913. It describes the 1,2-rearrangement of trityl amine derivatives to triaryl imines. It is comparable to a Beckmann rearrangement which also involves a substitution at a nitrogen atom through a carbon to nitrogen shift. As an example, triaryl hydroxylamines can undergo a Stieglitz rearrangement by dehydration and the shift of a phenyl group after activation with phosphorus pentachloride to yield the respective triaryl imine, a Schiff base.

<span class="mw-page-title-main">Deltic acid</span> Chemical compound

Deltic acid is a chemical substance with the chemical formula C3O(OH)2. It can be viewed as a ketone and double enol of cyclopropene. At room temperature, it is a stable white solid, soluble in diethyl ether, that decomposes between 140 °C and 180 °C, and reacts slowly with water.

<span class="mw-page-title-main">Oxazoline</span> Chemical compound

Oxazoline is a five-membered heterocyclic organic compound with the formula C3H5NO. It is the parent of a family of compounds called oxazolines, which contain non-hydrogenic substituents on carbon and/or nitrogen. Oxazolines are the unsaturated analogues of oxazolidines, and they are isomeric with isoxazolines, where the N and O are directly bonded. Two isomers of oxazoline are known, depending on the location of the double bond.

<span class="mw-page-title-main">Carbonyl reduction</span> Organic reduction of any carbonyl group by a reducing agent

In organic chemistry, carbonyl reduction is the conversion of any carbonyl group, usually to an alcohol. It is a common transformation that is practiced in many ways. Ketones, aldehydes, carboxylic acids, esters, amides, and acid halides - some of the most pervasive functional groups, -comprise carbonyl compounds. Carboxylic acids, esters, and acid halides can be reduced to either aldehydes or a step further to primary alcohols, depending on the strength of the reducing agent. Aldehydes and ketones can be reduced respectively to primary and secondary alcohols. In deoxygenation, the alcohol group can be further reduced and removed altogether by replacement with H.

<span class="mw-page-title-main">Rhodocene</span> Organometallic chemical compound

Rhodocene is a chemical compound with the formula [Rh(C5H5)2]. Each molecule contains an atom of rhodium bound between two planar aromatic systems of five carbon atoms known as cyclopentadienyl rings in a sandwich arrangement. It is an organometallic compound as it has (haptic) covalent rhodium–carbon bonds. The [Rh(C5H5)2] radical is found above 150 °C (302 °F) or when trapped by cooling to liquid nitrogen temperatures (−196 °C [−321 °F]). At room temperature, pairs of these radicals join via their cyclopentadienyl rings to form a dimer, a yellow solid.

<span class="mw-page-title-main">Trichloroacetonitrile</span> Chemical compound

Trichloroacetonitrile is an organic compound with the formula CCl3CN. It is a colourless liquid, although commercial samples often are brownish. It is used commercially as a precursor to the fungicide etridiazole. It is prepared by dehydration of trichloroacetamide. As a bifunctional compound, trichloroacetonitrile can react at both the trichloromethyl and the nitrile group. The electron-withdrawing effect of the trichloromethyl group activates the nitrile group for nucleophilic additions. The high reactivity makes trichloroacetonitrile a versatile reagent, but also causes its susceptibility towards hydrolysis.

<span class="mw-page-title-main">Tetramethylurea</span> Chemical compound

Tetramethylurea is the organic compound with the formula (Me2N)2CO. It is a substituted urea. This colorless liquid is used as an aprotic-polar solvent, especially for aromatic compounds and is used e. g. for Grignard reagents.

Hydroxylamine-<i>O</i>-sulfonic acid Chemical compound

Hydroxylamine-O-sulfonic acid (HOSA) or aminosulfuric acid is the inorganic compound with molecular formula H3NO4S that is formed by the sulfonation of hydroxylamine with oleum. It is a white, water-soluble and hygroscopic, solid, commonly represented by the condensed structural formula H2NOSO3H, though it actually exists as a zwitterion and thus is more accurately represented as +H3NOSO3. It is used as a reagent for the introduction of amine groups (–NH2), for the conversion of aldehydes into nitriles and alicyclic ketones into lactams (cyclic amides), and for the synthesis of variety of nitrogen-containing heterocycles.

References

  1. Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, ISBN   978-0-471-72091-1
  2. A.Aliab, C.Puzzarinid, "Cyclopropenyl cation – the simplest Huckel's aromatic molecule – and its cyclic methyl derivatives in Titan's upper atmosphere", Planetary and Space Science, Volume 87, October 2013, Pages 96-105. https://doi.org/10.1016/j.pss.2013.07.007
  3. De Meijere, A.; Faber, D.; Noltemeyer, M.; Boese, R.; Haumann, T.; Muller, T.; Bendikov, M.; Matzner, E.; Apeloig, Y. (1996). "Tris(trimethylsilyl)cyclopropenylium Cation: The First X-ray Structure Analysis of an α-Silyl-Substituted Carbocation". J. Org. Chem. 61 (24): 8564. doi:10.1021/jo960478e.
  4. Yadav, Arvind (2012). "Cyclopropenium Ion". Synlett. 23 (16): 2428–2429. doi: 10.1055/s-0032-1317230 .
  5. Ronald Breslow (1957). "Synthesis of the s-Triphenylcyclopropenyl Cation". J. Am. Chem. Soc. 79 (19): 5318. doi:10.1021/ja01576a067.
  6. Xu, Ruo; Breslow, Ronald (1997). "1,2,3-Triphenylcyclopropenium Bromide". Org. Synth. 74: 72. doi:10.15227/orgsyn.074.0072.
  7. Breslow, R.; Groves, J. T. (1970). "Cyclopropenyl Cation. Synthesis and Characterization". J. Am. Chem. Soc. 92 (4): 984–987. doi:10.1021/ja00707a040.
  8. Glück, C.; Poingée, V.; Schwager, H. (1987). "Improved Synthesis of 7,7-Difluorocyclopropabenzene". Synthesis. 1987 (3): 260–262. doi:10.1055/s-1987-27908.
  9. Buchholz, Herwig; Surya Prakash, G. K.; Deffieux, Denis; Olah, George (1999). "Electrochemical preparation of tris(tert-butyldimethylsilyl)cyclopropene and its hydride abstraction to tris(tert-butyldimethylsilyl)cyclopropenium tetrafluoroborate" (PDF). Proc. Natl. Acad. Sci. 96 (18): 10003–10005. Bibcode:1999PNAS...9610003B. doi: 10.1073/pnas.96.18.10003 . PMC   17831 . PMID   10468551.
  10. Bandar, Jeffrey S.; Lambert, Tristan H. (2013). "Aminocyclopropenium ions: synthesis, properties, and applications". Synthesis. 45 (10): 2485–2498. doi:10.1055/s-0033-1338516.
  11. Haley, Michael M.; Gilbertson, Robert D.; Weakley, Timothy J.D. (2000). "Preparation, X-ray Crystal Structures, and Reactivity of Alkynylcyclopropenylium Salts". Journal of Organic Chemistry. 65 (5): 1422–1430. doi:10.1021/jo9915372. PMID   10814104.
  12. 1 2 Hardee, David J.; Kovalchuke, Lyudmila; Lambert, Tristan H. (2010). "Nucleophilic Acyl Substitution via Aromatic Cation Activation of Carboxylic Acids: Rapid Generation of Acid Chlorides under Mild Conditions". Journal of the American Chemical Society. 132 (14): 5002–5003. doi:10.1021/ja101292a. PMID   20297823.
  13. Nogueira, J. M.; Nguyến, S. H.; Bennett, C. S. (2011). "Cyclopropenium Cation Promoted Dehydrative Glycosylations Using 2-Deoxy- and 2,6-Dideoxy-Sugar Donors". Journal of the American Chemical Society. 13 (11): 2184–2187. doi:10.1021/ol200726v. PMID   21548642.
  14. Yoshida, Zen'ichi; Yoneda, Shigeo; Hirai, Hideo (1981). "A Novel Synthesis of Pyrroles by the Reactions of Tris(alkylthio)cyclopropenium Salt with Amines". Heterocycles. 15 (2): 865. doi:10.3987/S-1981-02-0865 (inactive 2024-03-07).{{cite journal}}: CS1 maint: DOI inactive as of March 2024 (link)
  15. Chiang, T.; Kerber, R. C.; Kimball, S. D.; Lauher, J. W. (1979). "(η3-Triphenylcyclopropenyl) Tricarbonylcobalt". Inorganic Chemistry. 18 (6): 1687–1691. doi:10.1021/ic50196a058.
  16. Jiang, Yivan; Freyer, Jessica; Cotanda, Pepa; Brucks, Spencer; Killops, Kato; Bandar, Jeffrey; Torsitano, Christopher; Balsara, Nitash; Lambert, Tristan; Campos, Luis (2015). "The evolution of cyclopropenium ions into functional polyelectrolytes" (PDF). Nature Communications. 6 (1): 1–7. Bibcode:2015NatCo...6E5950J. doi: 10.1038/ncomms6950 . PMC   4354017 . PMID   25575214.